Вы находитесь на странице: 1из 33

This article was downloaded by: [Eindhoven Technical University]

On: 06 January 2015, At: 15:26


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Earthquake Engineering


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/ueqe20

A Detailed Dynamic Model of a Six-Axis


Shaking Table
a
A. R. Plummer
a
Centre for Power Transmission and Motion Control , University of
Bath , Bath, UK
Published online: 09 May 2008.

To cite this article: A. R. Plummer (2008) A Detailed Dynamic Model of a Six-Axis Shaking Table,
Journal of Earthquake Engineering, 12:4, 631-662, DOI: 10.1080/13632460701457264

To link to this article: http://dx.doi.org/10.1080/13632460701457264

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Journal of Earthquake Engineering, 12:631–662, 2008
Copyright © A.S. Elnashai & N.N. Ambraseys
ISSN: 1363-2469 print / 1559-808X online
DOI: 10.1080/13632460701457264

A Detailed Dynamic Model of a Six-Axis


1559-808X
1363-2469
UEQE
Journal of Earthquake Engineering
Engineering, Vol. 6, No. 1, June 2007: pp. 1–42

Shaking Table

A. R. PLUMMER
Detailed
A. R. Plummer
Dynamic Model of a Six-Axis Shaking Table

Centre for Power Transmission and Motion Control, University of Bath, Bath, UK
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

This article describes the modeling of a 5 m ´ 5 m 6 degree-of-freedom (DOF) shaking table and the
subsequent computer simulation of the system's dynamic characteristics. The simulation model is
required to aid the development of a closed-loop controller with a bandwidth specified up to 120 Hz,
and thus high order dynamics including valve response and structural effects are included. Signifi-
cant nonlinearities associated with the hydraulic and mechanical components are also included
(e.g., spool slew rate and saturation limits, valve overlap, manifold and valve body pressure losses,
friction, and geometric nonlinearities). Some model parameters are found from physical knowledge;
others are determined experimentally, and methods for estimating key parameters from experimen-
tal data are developed. The simulation is implemented using Simulink® and its multi-body mechani-
cal simulation tool SimMechanics®. The simulated response compares well with the measured
response of the table.

Keywords Servo-hydraulic Shaking Table; Seismic Testing; Hydraulic Modeling; Earthquake


Simulation; Parameter Estimation

1. Introduction
The shaking table remains an important tool for seismic testing. In the search for improved
earthquake simulation, increasing the accuracy with which a shaking table can replicate a
desired motion is of great interest. A thorough understanding of the dynamic characteris-
tics of a shaking table not only allows its limitations to be assessed, but also enables its
control algorithm to be optimised. This article concerns the modeling of a six DOF
(degree-of-freedom) table, and results in a set of high-order nonlinear equations suitable
for numerical solution. The model is derived for and verified against a particular example
shaking table, but the modeling techniques are applicable to a wide range of test systems.
Methods for the dynamic modeling of hydraulically actuated systems have been devel-
oped over a number of decades, [Merritt, 1967; Mcloy and Martin, 1980]. The derivation of
dynamic shaking table models of varying sophistication has been previously reported. The
simplest model of the single-axis positional response to a valve control signal input is a
series gain and integrator. The addition of an empirical first-order lead/lag, identified from
experimental data, is reported in Chase et al., [2005]. More commonly, however, a second-
order lag term is added to model the table mass interacting with the hydraulic “oil-column”
compressibility, e.g., Shimizu et al., [2002]. In some cases, it has been recognized that the
valve response has a significant impact on system dynamics, and this has been modeled
along with the integrator and second-order hydraulic resonance term. The valve response
model used has variously been a delay [Trombetti and Conte, 2002], a first-order lag

Received 21 December 2006; accepted 14 May 2007.


Address correspondence to A. R. Plummer, Centre for Power Transmission and Motion Control, University
of Bath, Bath BA2 7AY, UK; E-mail: A.R.Plummer@bath.ac.uk

631
632 A. R. Plummer

[Kuehn et al., 1999], and a second-order lag [Iwasaki et al., 2005; Kakegawa et al., 2003;
Sato et al. 2002]. The likelihood that structural compliance in the test rig will affect perfor-
mance has not often been considered, however, reaction mass dynamics have been mod-
eled in one case [Conte and Trombetti, 2000; Trombetti and Conte, 2002].
Many studies have considered uni-axial shaking tables, or just one axis of multi-axial
systems. A number of linear multi-variable models have also been reported. The three DOF
planar model of Shimizu et al. [2004] includes a linear relationship between actuator and
table co-ordinates. Linear modeling of Japan’s E-defense shaking table has enabled the six
hydraulic modal resonances to be predicted [Kakegawa et al., 2003; Ogawa et al., 2001].
Most models have been linear. However, the valve saturation non-linearity, resulting
in a limit on actuator velocity has been included in Chase et al., [2005]. The need for
including nonlinear characteristics in test rig models has been noted, such as geometric
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

effects, fiction, and variable hydraulic stiffness [Shortreed et al., 2002].


Seismic testing of large structures presents a significant control challenge, and is the
subject of academic research as well as the application of advanced industrial control tech-
niques [Plummer, 2007]. Shaking tables for testing large structures at full scale have high
mass and long-stroke actuators, giving low hydraulic resonant frequencies; for example, the
E-defense table with a 1,200 tonne payload has resonant frequencies in the range 3–8 Hz, yet
accurate control is required up to 15 Hz [Ogawa et al., 2001]. Alternatively, a scale model of
the test structure can be used, but this necessitates compressing the timescale of the earth-
quake signal and hence demands a higher bandwidth for the closed-loop system. For the
table considered in this article, hydraulic resonances are typically in the range 20–30 Hz, but
test signals at up to 120 Hz must be followed. The need to test at frequencies beyond the
hydraulic resonance clearly means that modeling and compensation for this characteristic is
essential. The servovalve response, particularly its phase lag, will also have a significant
impact, and a high-performance controller cannot be designed without accounting for it.
The model developed in this article is intended for accurate simulation of table
dynamics. The model has been used in simplified form for controller design, and in its full
form for predicting controller performance in simulation, but controller development is
not covered in this article. The relatively straightforward task of modeling a known con-
troller is not described, nor is specimen modeling included.
Section 3 contains the analytical model development, and Sec. 4 describes the very
important task of determining accurate model parameters. The implementation of the
model using Simulink is outlined in Sec. 5, and results are presented in Sec. 6. Appendix 1
details some parameter estimation methods, and Appendix 2 summarizes the notation used
in the article.

2. IWHR Shaking Table


The table modeled in this article is at IWHR, Beijing (Fig. 1). It is a six DOF system
driven by seven actuators. The table has a 5 m × 5 m surface area and a payload capacity
of 20 tonnes. The three horizontal actuators drive the table through pushrods; the cylinder
bodies themselves are fixed. These actuators are equipped with 3-stage servovalves. The
four vertical actuators are jointed at both ends, and are each equipped with four 2-stage
servovalves. Figure 2 helps to clarify the mechanical arrangement. Ball joints are used
throughout. The blind end of the double-ended vertical actuators are pressurised to support
the dead weight of the table.
Horizontal velocities of nearly 0.5 m/s can be achieved, and accelerations of over
10m/s2. The intended test frequency range for this table extends up to 120 Hz, functioning
either in position or acceleration control.
Detailed Dynamic Model of a Six-Axis Shaking Table 633

2Lh
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

2Lv 2Lv

FIGURE 1 IWHR shaker table.

Specimen

FIGURE 2 IWHR shaker table schematic, showing actuator connections.

3. Modeling

3.1. Actuator Model


The dynamic characteristics of a hydraulic cylinder and servovalve are modeled in this Sec-
tion. The model derivation largely follows established procedures [Merrit,1967; McLoy
and Martin, 1980], with some variation in detail which has been found useful in practice,
for example the nature of the valve dynamics model. All aspects of the behavior which can
have a significant impact on the overall system response are included. These are:
• fluid compressibility;
• variable cylinder oil volumes (giving variable stiffness);
• internal cylinder leakage;
• cylinder cross-port bleed (an orifice which is adjusted to increase damping);
634 A. R. Plummer

• valve orifice pressure-flow characteristic;


• valve overlap;
• valve body pressure drop;
• manifold pressure drop and oil volume;
• valve spool dynamics;
• maximum valve opening;
• valve spool slew rate limit.
Note that the preload system included in the vertical actuators for supporting the weight of
the table and specimen is not modeled in this article.
Figure 3 shows the construction of one of the hydraulic actuators. It is has a double-
acting, equal-area cylinder with built in position measurement and cylinder-mounted
valves. The cylinder flow equations are
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

v1
q1 = Ay&c + p&1 + Cl ( p1 − p2 ) + Cb f ( p1 − p2 ) (1)
B

v2
q2 = Ay&c − p& 2 + Cl ( p1 − p2 ) + Cbφ ( p1 − p2 ) (2)
B

where the function f(•) is a square root with modified sign:

φ ( p) = sgn( p) p (3)

and where B is the fluid bulk modulus, Cl is a cross-piston leakage coefficient, and Cb is a
cross-port bleed coefficient. The function sgn(p) is + 1 for p ≥ 0, and −1 for p < 0. The
leakage is assumed to be laminar flow, and so flow and pressure drop are directly propor-
tional, but the cross-port bleed is assumed to be turbulent flow. The total oil volumes
trapped between the valve orifices and the piston, on either side of the cylinder, are:

v1 = V + Ayc (4)

v2 = V − Ayc (5)

Servovalve Cross-port bleed

Manifold
Position measurement (yc) q1 q2

yc
A A
Pressure p1 Pressure p2
Leakage Volume v2
Volume v1

FIGURE 3 A double-ended hydraulic actuator.


Detailed Dynamic Model of a Six-Axis Shaking Table 635

where V is the volume when the piston is at mid-stroke, which is defined as the zero posi-
tion for yc.
The hydraulic force is defined as:

fh = ( p1 − p2 ) A. (6)

At the end of its stroke the piston enters a hydraulic buffer to cushion the impact; it is
useful to include the buffer in a simulation model, but as this is beyond the displacement
range in which a meaningful test can be performed, it is not described here.
The relationship between valve control signal, pressure and flow is critical to the
behavior of the actuator. As with most large shaking tables, the horizontal actuators in this
example are fitted with three-stage servovalves. The typical construction of a three-stage
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

valve is shown in Fig. 4. The first stage consists of a torque motor controlling flow via a
nozzle-flapper arrangement, the second stage is a spool valve with mechanical feedback to
the first stage (the feedback spring), and the third stage is also a spool valve but with elec-
tronic closed-loop control of spool position. The normalized third-stage spool displace-
ment is denoted by x (i.e., x = ±1 is the maximum displacement). The actual orifice
opening is denoted xv. Although the valve would nominally be zero-lapped, slight overlap
is typically present due to manufacturing tolerances. The following empirical model is
used to account for overlap (see Fig. 5):

xv = x / 2 for x ≤ L (7)

Control signal, u (3rd stage


spool position command) 3rd stage spool position: x
Electronic closed-loop
controller
Permanent magnet

Flapper N N Torque motor


Feedback
spring S S
2nd stage spool

x
pv1 pv2

pa q1 Pr q2 pa
3rd stage spool To cylinder (p1) From cylinder (p2) Position sensor

FIGURE 4 A three-stage valve.


636 A. R. Plummer

Effective orifice
opening, xv

L /2 +1
–1 0 L Spool displacement, x

Overlap region
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

FIGURE 5 Empirical overlap model.

xv = x − L / 2 for x > L (8)

xv = x + L / 2 for x < − L. (9)

The third-stage spool orifice equations are

q1 = K v xv φ ( pa − pv1 ) for xv ≥ 0 (10)

q1 = K v xv φ ( pv1 − Pr ) for xv < 0 (11)

q2 = K v xv φ ( pv 2 − Pr ) for xv ≥ 0 (12)

q2 = K v xv φ ( pa − pv 2 ) for xv < 0 (13)

where Kv is the valve flow coefficient. The pressures pv1 and pv2 are within the valve, adja-
cent to the 3rd stage spool; they differ from the cylinder pressures p1 and p2 due to pressure
losses in the manifold and valve body. Similar pressure losses exist on the supply side, i.e.,
between the high pressure line from the hydraulic power unit (pa) to the valve spool, and
from the valve spool to the return line (Pr). However, a lumped parameter model will be
used, where all these losses are represented by a single manifold (Kf) and valve body (Kb)
flow coefficient, assumed to affect only the cylinder pressures. Therefore,

⎛ 1 1 ⎞
pv1 − p1 = sgn(q1 ) ⎜ 2 + 2 ⎟ q12 (14)
⎜⎝ K b K f ⎟⎠

⎛ 1 1 ⎞
p2 − pv 2 = sgn(q2 ) ⎜ 2 + 2 ⎟ q2 2 . (15)
⎜⎝ K b K f ⎟⎠
Detailed Dynamic Model of a Six-Axis Shaking Table 637

The return pressure Pr is assumed to be constant. The design supply pressure Ps is also
assumed to be constant, but if the system flow requirement is greater than the pump flow,
the accumulators discharge and the pressure of oil supplied to the valve (pa) reduces (see
Sec. 3.2). These pressures are shown on the circuit diagram in Fig. 6.
An empirical model is used for the valve dynamics. The model is linear for the usual
operating range of the valve, such that the normalized spool displacement x is related to
the normalized control signal u by

x = V (s)u (16)

where V(s) is a second-order transfer function plus delay:


Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

e − sD
V ( s) = .
1 2 2ζ v (17)
s + s +1
ω nv 2 ω nv

This model has been used successfully in the past (see, for example, Plummer, 2001).
However, two important nonlinearities should not be neglected. Firstly there is the maxi-
mum spool displacement so that −1 ≤ x ≤ 1. Also, there is a maximum flow which can be
delivered by the second stage of the valve, giving a maximum velocity for the third-stage
spool; this is called the slew rate limit (Sr). Both these are included in the model by imple-
menting Eq. (16) as shown in the block diagram of Fig. 7. If the constraints of blocks (a)
and (b) are not contravened then this implementation reduces back to Eq. (16).

3.2. Accumulator Model


Typically, small accumulators are mounted close to each actuator to accommodate short-
duration, high-flow requirements. A larger accumulator bank is mounted close to the
pump or substation, often supplying oil to the whole system throughout a high-flow test.
The small local accumulators are neglected in this model.

Main
accumulator Local
To other axes accumulator

max
qp qsi
Power pack
(max
pressure Ps)
Pr Pa

Servovalve

Variable restrictor
(cross-port bleed)
Cylinder

FIGURE 6 Circuit diagram for one actuator.


638 A. R. Plummer

(a) (b)
.
ωnv 2 sr
v x = 0 for x ≥ 1 .
x x
u −sD
+ –sr 1
e s + 2ζvωnv sr & sgn(v) = +1 s
–sr .
– x = 0 for x ≤ –1
& sgn(v) = –1
.
otherwise x = v

FIGURE 7 Valve dynamics model.

The supply line flow into the valve of actuator i is given by:
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

qsi = q1 for x ≥ 0 (18)

qsi = −q2 for x < 0. (19)

A polytropic relationship between pressure and accumulator gas volume is usually used:

n
⎡V ⎤
pa = Ps ⎢ s ⎥ (20)
⎣ va ⎦

where Ps is the design supply pressure for the system (the hydraulic power unit relief
valve cracking pressure), Vs is the gas volume at that pressure, and va is the actual gas vol-
ume at current pressure pa. A reasonable average value for the polytropic index is n = 1.6.
Volume Vs can be determined from the accumulator volume and its precharge pressure (V0
and P0) by assuming an isothermal compression

P0V0
Vs = . (21)
Ps

When the sum of flows required by all actuators exceeds the available pump flow Qp, the
integral of the excess flow is the oil volume removed from the accumulator, thus

⎛ ⎞
va = Vs + ∫ ⎜ ∑ qsi − q p ⎟ dt (22)
⎝ i ⎠

where

q p = ∑ qs i for ∑ qsi < Qp and va ≤ Vs (23)


i i

q p = Q p otherwise. (24)

The available pump flow, Qp, is the actual pump flow less the flow required to accommo-
date valve leakage, hydrostatic bearings, etc.
Detailed Dynamic Model of a Six-Axis Shaking Table 639

Note that tests presented later in this article do not discharge the accumulator and thus
this part of the model is not validated in the current work.

3.3. Mechanical Model


Equations (1–24) allow the hydraulic force to be determined for any valve control signal
and piston-to-cylinder relative motion. To complete the model, the behavior of the
mechanical system needs to be added. A lumped parameter model is used, consisting of a
collection of rigid bodies connected by joints, with compliance included at the cylinder
mounting and piston rod-to-table connection. For the purpose of exposition, the specimen
is considered as another rigid body attached to the table; clearly, a more complex speci-
men model should be used where necessary. For the table under consideration, the hori-
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

zontal cylinders are fixed, and linked to the table by pushrods; the vertical cylinder bodies
are moving, jointed to ground at their lower end. A horizontal actuator, including the com-
pliance modeling, is illustrated in Fig. 8. Table 1 describes the parameters required for the
mechanical model. Note that the table structure itself has been designed to be very stiff

Mc Mp
Km fh Kp

Table and
Specimen
Cm Cp (mass Mt)
Ym Yc Yp
Ya
Yt

FIGURE 8 Horizontal actuator, included structural compliance.

TABLE 1 Mechanical model parameters


Component Parameter Component Parameter
All actuators Horizontal actuators only
Table ball joint Coordinates (x,y,z) Pushrod (table end) Mass
Friction torque Centroid (x,y,z)
Ground ball joint Coordinates (x,y,z) Moments of inertia
Friction torque Pushrod (piston end) Mass
Cylinder mounting Stiffness (Km) Centroid (x,y,z)
Damping (Cm) Moments of inertia
Table connection Stiffness (Kp) Vertical actuators only
Damping (Cp) Cylinder body Mass (Mc)
Piston rod plus ball Mass (Mp) Centroid (x,y,z)
Table/specimen Moments of inertia
Bare table Mass Centroid (x,y,z)
Centroid (x,y,z) Piston rod plus ball Moments of inertia
Moments of inertia
Specimen Mass
Centroid (x,y,z)
Moments of inertia
640 A. R. Plummer

and does not possess any flexural modes in the frequency range of interest. Likewise,
other than cylinder mounting stiffness, there is no evidence of dynamics associated with
the reaction mass upon which the system is mounted.
The equations of motion are not included here. They are generated and solved auto-
matically by Multi-Body Simulation (MBS) software. In this case, SimMechanics, a tool-
box of Simulink, is used. This generates Newton’s equations of motion from a block
diagram system representation, and one of a range of differential equation solvers can be
chosen to integrate forces and torques to find the resultant motion. As well as the dynamic
characteristics, this will accurately solve the mechanism kinematics so that large stroke
motions are realistically simulated.
Note that in some actuators a friction force might be significant, but in this case actu-
ators with hydrostatic bearings and no internal seals are used, and so the friction force is
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

very low and can be neglected. However, the ball joints do exhibit noticeable friction. This
is modeled as a friction torque which opposes the resultant angular velocity experienced
by the joint. A constant friction torque is used, which is a good assumption when the force
transmitted by the joint is small, as the normal load on the ball is dominated by the joint
preload. However, an improvement would be to increase the friction torque as a function
of joint force when the force is high.

3.4. Linear Mechanical Model


A linear approximation for the mechanical model can be derived. Such a model is useful
for analytical studies and controller design. Due to relatively short actuator strokes, the
geometric nonlinearities are fairly small, and the linear model will normally be a good
approximation.
The lengths Ya, Yc, Ym, Yp, and Yt shown in Fig. 8 are for the steady-state zero force
condition with zero actuator displacement, and changes in these lengths will be denoted
ya, yc, ym, yp, and yt, respectively. The displacement of the cylinder body on its mounting is
given by

1
ym = − fh (25)
M c s + Cm s + K m
2

and the displacement of the piston rod relative to the table is

1
ya − yt = ft (26)
Cps + K p

where ft is the force exerted on the table by this actuator, given by

ft = fh − M p s2 ya . (27)

Combining (26) and (27):

Cps + K p ⎛ 1 ⎞
ya = ⎜ fh + yt ⎟ . (28)
M ps + Cps + K p ⎝ Cps + K p
2

Detailed Dynamic Model of a Six-Axis Shaking Table 641

Thus, the relative piston-to-cylinder displacement can be deduced from

yc = ya − ym (29)

along with Eqs. (28) and (25).


All that remains is to determine the table motion as a result of the forces acting
upon it. A linearized six DOF model is developed below, following the approach of
Plummer [2004]. A vector of Cartesian forces and moments acting on the table is
given by

T
f%t = ⎡⎣ f x fy fz mx my mz ⎤⎦ (30)
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

and the vector of N actuator forces is

ft = [ ft1 ft 2 ft 3 ... ftN ]T . (31)

Similarly, define Cartesian and actuator displacement vectors as

T
y% t = ⎡⎣ y x yy yz jx jy j z ⎤⎦ . (32)

y t = [ yt1 yt 2 yt 3 ... ytN ]T . (33)

The actuator force vector is transformed to the Cartesian force vector using a Jacobian
matrix J:

f%t = J T ft . (34)

In Cartesian space, a mass matrix M is used to perform the linear transformation from
table force to acceleration

&&
y% t = Mf%t . (35)

Table motion is transformed back from Cartesian space to actuator space thus,

y t = J y% t . (36)

Combining these transformations gives

1
yt = 2
JMJ T ft . (37)
s

For the example table in question, actuators are symmetrically disposed around the centre-
line with the spacing indicated in Fig. 1, giving the following Jacobean:
642 A. R. Plummer

⎡0 0 1 − Lv Lv 0 ⎤
⎢0 0 1 − Lv − Lv 0 ⎥⎥

⎢0 0 1 Lv − Lv 0 ⎥
⎢ ⎥ (38)
J=⎢0 0 1 Lv Lv 0 ⎥.
⎢ −1 0 0 0 0 − Lh ⎥
⎢ ⎥
⎢ −1 0 0 0 0 Lh ⎥
⎢⎢ 0 0 ⎥⎥⎦
⎣ 1 0 0 0

And the simplest mass matrix would be where the combined table/specimen centroid is at
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

the point where the lines of actuator symmetry in all planes meet, giving:

⎡ Mt 0 0 0 0 0⎤
⎢ 0 Mt 0 0 0 0 ⎥⎥

⎢ 0 0 Mt 0 0 0⎥
M=⎢ 0 0 0 I xx I xy

I xz ⎥ . (39)

⎢ 0 0 0 I yx I yy I yz ⎥
⎢ ⎥
⎢ 0 0 0 I zx I zy I zz ⎥⎦

The complete model for one actuator, incorporating the linear mechanical model as an
example, is illustrated in Fig. 9. This model needs to be replicated for each actuator, shar-
ing the accumulator and table inertia equations, to form the complete system model.

4. Parameterization

4.1. Parameterization Methods


Model parameter values have been determined, partly from a priori known physical infor-
mation, and partly by estimation from experimental data. In a few instances, typical values
have been used where precise values are not known and cannot easily be estimated. In
some cases estimated values can be checked against physical design data. Table 2 gives
parameters, values, and how they have been found, using one horizontal actuator as an
example (actuator 5).

4.2. Estimation of Valve Dynamics


A measurement of spool position is available, and so the transfer function V(s) between
control signal and spool displacement can be determined. Spectral analysis of data col-
lected with a random (pink noise) actuator command signal gives the valve frequency
response shown in Fig. 10.
A least squares fit of V(s) to the measured response in the frequency domain gives the
parameters:

w n v = 754 rad/s zv = 0.70 D = 0.002s


Detailed Dynamic Model of a Six-Axis Shaking Table 643

p1 fh
Valve body & manifold
pressure drops p2 Eqn 6
Eqn 14 – 15
pv2 pv1

u Valve x Overlap xv q1
dynamics Eqn 7 – 9
Valve q2
(Fig. 7)
orifices
Eqn 10 – 13
pa

Accumulator qsi Supply flow Cylinder flow


Eqn 20 – 24 Eqn 18 – 19 Eqn 1 – 2
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Other
axes v1
Oil
volumes v2
Eqn 4 – 5
yc

(a) Hydraulic model

Mounting ym
yc
compliance Eqn 29
Eqn 25

fh
Piston mass
ya
Eqn 27

ft Pushrod compliance
Eqn 26
yt

Table plus specimen


Other inertia Other
axes (6 DOF) axes
Eqn 37

(b) Mechanical model (linear version used as an example)

FIGURE 9 Complete model for one actuator.

4.3. Estimation of Cylinder Mounting and Pushrod Stiffness/Damping


As shown in Appendix 1, a relationship between piston-to-cylinder displacement yc and
absolute piston acceleration ÿa can be derived as:

⎛ Cp K p s + 1⎞ K p 2 Cp Kp 2
s2 (s2 yc − &&
ya ) = ⎜ ⎟ ya −
s && ya ) −
s(s2 yc − && (s yc − &&
ya ). (40)
⎝ C m K m s + 1⎠ K M Mt Mt
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

644
TABLE 2 Main parameters for table and example actuator (actuator 5)
Parameter Value Comment
Parameter values from physical information / design data
Table mass Mt 23 tonnes
Table centroid co-ords. [0, 0, −0.25] m Origin at center of table top
Table moments of inertia [71 48 97]*1000 kgm2
Piston area A 127 cm2
Cylinder/manifold volume V 700 cm3 One side of piston
Cylinder leakage Cl 11.9 × 10−12 m5/s/N From design value of 5 l/min at 70 bar
Supply pressure Ps 275 bar Allowing for small pipe pressure drop
Return pressure Pr 5 bar Allowing for small pipe pressure drop
Table joint coordinates (x,y,z) 1–4 [±1.75, ±1.75, −1.31] Origin at centre of table top (Lv = 1.75 m, Lh = 2.8 m)
5,6 [0.45, ±2.8, −0.25] m
7 [0, −2.8, −0.25] m
Ground joint coordinates (x,y,z) 1–5 [±1.75, ±1.75, −2.4] Origin at center of table top
5,6 [1.75, ±2.8, −0.25] m
7 [0, −3.93, −0.25] m
Valve slew rate limit Sr 700 /s
Valve body flow coefficient Kb 3.78 × 10−6 m4/s/N½ Equivalent to 600 l/min @ 70 bar
Manifold flow coefficient Kf 2.52 × 10−6 m4/s/N½ Equivalent to 400 l/min @ 70 bar
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Experimental parameter values


Mounting stiffness Km 1750 kN/mm Estimated as shown in Sec. 4.3
Mounting damping Cm 987 kNs/m
Pushrod Stiffness Kp 563 kN/mm
Pushrod damping Cp 318 kNs/m
Estimated as shown in Sec. 4.4
Bulk modulus B 1.55 GN/m2 From estimated hydraulic stiffness of Kh = 715
kN/mm at mid-stroke.
Cross-port bleed Cb 3.86 × 10−8 m4/s/N½ From estimated series damping of Ch = 8410 kNs/m.
Equivalent to 6.1 l/min @ 70 bar
Valve flow co-efficient Kv 1.75 × 10−6 m4/s/N½ Equivalent to 196 l/min @ 35 bar per orifice, c.f.
rated valve flow of 200 l/min @ 35 bar per orifice.
Valve dead time D 2 ms Estimated as shown in Sec. 4.2
Valve natural freq wnv 754 rad/s (120 Hz)
Valve damping ratio zv 0.7
Valve overlap L 0.7% Manually adjusted to fit measured behavior
at turn-around.
Friction torque, joints 1–4 450 Nm
Friction torque, joints 5–6 2600 Nm
Friction torque, joint 7 560 Nm

645
646 A. R. Plummer

Amplitude Ratio (dB)


Measured
0

Delay model
–5 Estimated
2nd V(s)
order lag + delay
Measured
–10
100 101 102
Frequency (Hz)

50
0
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Measured
Phase (deg)

–50
–100
–150
–200
–250 0
10 101 102
Frequency (Hz)

FIGURE 10 Valve frequency response: spool position x over control signal u (actuator 5).

This relationship includes the table mass Mt (known), and the structural stiffness and
damping values associated with the cylinder mounting and the joints/pushrod. A method
for estimating these stiffness and damping values from the yc to ÿa frequency response is
also shown in the Appendix. This frequency response, calculated from data acquired with
a pink noise actuator command signal, is shown in Fig. 11, together with the least squares
model fit. This gives the values:

K m = 1750 kN/mm, Cm = 987 kNs/m, K p = 563 kN/mm, C p = 318 kNs/m.

4.4. Estimation of Hydraulic Parameters


Also shown in Appendix 1, an approximate linear relationship between piston-to-cylinder
displacement yc and absolute piston acceleration ÿa can be derived as (Eq. (A32)):

Kv Ps − Pr V (s) ⎛ 1 1 ⎞
yc = u+⎜ + F (s) yc (41)
A 2 s ⎝ K h sCh ⎟⎠

where F(s) is the known transfer function between displacement yc and force on the table.
The actuator is modeled as a stiffness Kh in series with a damper Ch, and these two param-
eters along with the valve flow coefficient Kv can be estimated from the control signal (u)
to measured displacement (yc) frequency response. The response for actuator 5 and the
least squares fit are shown in Fig. 12. This gives values

K v = 1.75 × 10 −6 m 4 /s/N1 2 , K h = 715 kN/mm, Ch = 8410 kNs/m.


Detailed Dynamic Model of a Six-Axis Shaking Table 647

120

Amplitude ratio (dB) 100

80

60

40
100 101 102

200
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

150
Phase (°)

100 Measured
Linear model
50

0
100 101 102
Frequency (Hz)

FIGURE 11 Frequency response of piston acceleration ÿa over piston-to-cylinder relative


position yc (actuator 5).

–20
Amplitude ratio (dB)

–40

–60

–80
100 101 102

–50

–100
Phase (°)

–150
Measured
–200
Linear model
–250

–300
100 101 102
Frequency (Hz)

FIGURE 12 Frequency response of piston-to-cylinder position yc over control signal u


(actuator 5).
648 A. R. Plummer

The actuator stiffness value, given that the data were collected with the piston near mid-
stroke, allows the effective fluid bulk modulus to be deduced as per Eq. (A23) (using
known piston area and oil volumes). The series damping value accounts for both the cross-
port bleed and the cylinder internal leakage. The cylinder leakage flow is 5 l/min at a 70
bar pressure differential according to design data. Equation (A24) can be used to deter-
mine the cross-port bleed coefficient; taking the typical operating pressure drop ΔP to be
70 bar, the cross-port bleed flow co-efficient is as shown in Table 2.

5. Simulation Model
The model equations developed in Sec. 3 have been used to create a Simulink simulation
model for the seismic table. Figure 13 is the top level Simulink block diagram. The ‘Actu-
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

ators’ block includes the hydraulic system model; within it are seven individual actuator
models of the form shown in Fig. 14. The mechanical model is implemented using
Simulink’s SimMechanics toolbox. This model, which is within the ‘Mechanical model’
block in Fig. 13, is shown in Fig. 15. A simple visualisation of the mechanical model is
produced by SimMechanics (Fig. 16).

yc
LVDTs 1
Cylinder LVDTs
Actuator vel

Actuator accel 8
Piston accelerometers
Load cells
u fh
1 Valve drive Force Hyd forces Terminator
Control signals Table position 4
Position DeltaP 2 Table position (6DOF)
DeltaPs Table acceleration 5
Velocity PreloadP 7 Table acceleration (6DOF)
PreloadPs Table load 6
Actuators
Specimen load (6DOF)
accelerometers 3
Table accelerometers
SIMMECHANICS
Mechanical model

FIGURE 13 Top level Simulink block diagram.

P1act
P1valve
P2act

Q1
P2valve
Q2

Manifold

P1
Q1 Q1 fh
Force 1
P2
u force
1 U Q2 Q2
Valve drive P1
4 Ps 2
X 2 Position
Accumulator pressure Pr DeltaP
Position
-C- Valve P2
3 Velocity
Return Pressure Velocity
Cylinder

FIGURE 14 Simulink actuator model.


Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Conn1
5 Conn2
table load 7 Rotmatrpy Conn3
Table position Conn4
Table load to angles(deg) 6 Conn5 accel 8
Conn6 accelerometers
Specimen Table acceleration Conn7
Body Sensor
Conn8
Conn9
Accelerometers

CS9
CS10
CS11
CS12
CS13
CS14
CS15
CS16
CS17
CS18
CS19
Table

CS1
CS2
CS3
CS4
CS5
CS6
CS7
[D1] [D2] [D3] [D4]

Pushrod 5 Pushrod 6 Pushrod 7

Top
Top
Top
Top
Base Top
Base Top
Base Top

Moving Moving Moving Moving


Cylinder 1 Cylinder 2 Cylinder 3 Cylinder 4
[D5] [D6] [D7]

Outputs Force
Base
Outputs Force
Base
Outputs Force
Base
Outputs Force
Base
[l1] [l2] [l3] [l4]

Top
Top
Top

Fixed Fixed Fixed


Cylinder 5 Cylinder 6 Cylinder 7

Outputs Force
Base
Outputs Force
Base
Outputs Force
Base

Env VERTICAL ACTUATORS [l5] [l6] [l7]


Machine
Environment [l1] leg1
[l2] leg2 LVDTs 1
[l3] leg3 LVDTs
[D1]
[D2] [l4] leg4 Actuator vel 2
[D3] Actuator vel HORIZONTAL ACTUATORS
[D4]
[l5] leg5
1 [D5] [l6] leg6 Actuator accel 3
Hyd forces [D6] Actuator accel
[D7]
[l7] leg7
[D8] [l8] leg8 Actuator load 4
Output configuration Load cells

FIGURE 15 SimMechanics mechanical model.

649
650 A. R. Plummer

0
Z-axis

–1

–2
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

–3

–4

–2
2
0 1
0
–1
2 –2
X-axis –3
4 –4
Y-axis

FIGURE 16 SimMechanics mechanical model visualization.

The numerical integration of the hydraulic and mechanical equations is performed by


an explicit Runge-Kutta (4,5) solver (Simulink’s ODE45) with variable step size. The esti-
mated error of each state is constrained to be within 0.1% at all time steps, and the speci-
fied maximum step size is 0.1 ms. A larger maximum step size was found to compromise
the stability of the numerical integration. The simulation speed is around 2% of real-time
using a 2 GHz PC.

6. Results
The simulation model can be excited in a similar manner to the real table, and frequency
responses determined from the acquired data. Figures 17 and 18 mirror the responses of
Figs. 11 and 12, except that now the simulated responses are compared with the measured
ones. Figure 17 shows a very close match, which is unsurprising as this relationship
between measured piston position and acceleration is not affected by most of the system
nonlinearities. Figure 18, which is dependent on the characteristics of the whole hydraulic
actuator system, shows some discrepancy in phase but overall the match is still reason-
able. Other linear and angular directions show a similarly good match between simulated
and measured responses.
The nonlinear characteristics can be explored more fully by inspecting time
responses. Figures 19, 20, and 21 show measured responses to a 1.7 Hz sinusoidal posi-
tion command signal in the x-direction of different amplitudes. The measured control sig-
nal is used to drive the simulation, and the resulting simulated responses are plotted for
comparison. The control signal for the 40 mm amplitude response (Fig. 21) peaks at
nearly 100%.
Detailed Dynamic Model of a Six-Axis Shaking Table 651

120

Amplitude Ratio (dB)


100

80

60

40
100 101 102
Frequency (Hz)
–100
–150
Phase (deg)
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

–200
–250 Simulation
Measured
–300
–350
100 101 102
Frequency (Hz)

FIGURE 17 Frequency response of piston acceleration ÿa over piston-to-cylinder relative


position yc (actuator 5).

–20
Amplitude Ratio (dB)

–40

–60 Simulation
Measured
–80
100 101 102
Frequency (Hz)

–50

–100
Phase (deg)

–150

–200

–250

–300
100 101 102
Frequency (Hz)

FIGURE 18 Frequency response of piston-to-cylinder position yc over control signal u


(actuator 5).

The gross simulated amplitudes of all signals are very accurate. A number of parame-
ters significantly affect this match; for example, if manifold and valve body pressure
losses were not included the simulation would significantly over-estimate the motion
amplitude when the control signal is large.
652 A. R. Plummer

4 0.6
Simulation

Piston acceleration (m/s2)


LVDT position (mm)
Experiment 0.4
2 0.2
0
0
–0.2

–2 –0.4
–0.6
–4 –0.8
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

10 0.6
Differential pressure (bar)

Table acceleration (m/s2)


0.4
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

5 0.2
0
0
–0.2
–0.4
–5
–0.6
–10 –0.8
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

FIGURE 19 Actuator 5 (x-direction): 4 mm amplitude sinusoid.

30 3
Piston acceleration (m/s2)

Simulation
2
LVDT position (mm)

20 Experiment

1
10
0
0
–1
–10 –2

–20 –3
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

30 3
Differential pressure (bar)

Table acceleration (m/s2)

20 2

10 1

0 0

–10 –1

–20 –2

–30 –3
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

FIGURE 20 Actuator 5 (x-direction): 20 mm amplitude sinusoid.

The precise acceleration and cylinder pressure transients that occur when the valve
passes through its null position and the direction of travel reverses (sometimes called turn-
around knock) are notoriously difficult to model. However, the form of the transients in
Detailed Dynamic Model of a Six-Axis Shaking Table 653

40 5
Simulation

Piston acceleration (m/s2)


Experiment
LVDT position (mm)

20

0 0

–20

–40 –5
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

50 5
Differential pressure (bar)

Table acceleration (m/s2)


Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

0 0

–50 –5
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

FIGURE 21 Actuator 5 (x-direction): 40 mm amplitude sinusoid.

the measured and simulated responses are similar. The transients are influenced by fric-
tion, which particularly affects the cylinder differential pressure signal at low amplitudes.
They are also influenced by the valve overlap model.
The match between measured and simulated responses in other linear and angular
directions is similar to that presented here. The largest discrepancy occurs in the vertical
(z-direction) response at low amplitude (Fig. 22). A discrepancy in cylinder differential
pressure is seen, which can be attributed to bending of the pushrods; in the model the
pushrods only have axial compliance.

7. Conclusions and Discussion


A detailed model of a six DOF servohydraulic shaking table has been derived. Simulink is
used for numerical solution of the model equations, together with its multi-body simula-
tion toolbox SimMechanics. The resulting prediction of dynamic characteristics is a good
match to the measured response. This is achieved by paying careful attention to the model-
ing of features such as structural compliances and servovalve response. It is also shown
how a number of key parameters can be estimated from experimental data, reducing the
reliance on design information. Nevertheless, a good knowledge of physical design
parameters is helpful, and in some cases still essential.
The table simulation is verified without a specimen. Both the full simulation model
and the linear approximation allow for adding a rigid body specimen model in the deriva-
tion. Adding the dynamics of a complex specimen would be relatively straightforward
using Simulink/SimMechanics if the dynamic characteristics are known. Given that the
bare-table model has been carefully verified, it would be expected that the effect of the
specimen dynamics on the table response would be accurately predicted by the
simulation.
654 A. R. Plummer

3 0.6
Simulation

Piston acceleration (m/s2)


2 Experiment
LVDT position (mm)
0.4
1
0.2
0
0
–1

–2 –0.2

–3 –0.4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

10 0.6
Differential pressure (bar)

Table acceleration (m/s2)


Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

0.4
5
0.2
0
0
–5
–0.2

–10 –0.4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Time (s) Time (s)

FIGURE 22 z-direction: 3 mm amplitude sinusoid.

Although not described in this article, the main use of the model in this case has
been in controller development. A simplified form of the model was used to help design
the controller, and the full model was used to test the controller in simulation. Subse-
quently, the real controller was tested against the simulation running in real-time. For
this purpose Simulink’s real-time code generator was used to create code to run on a
dSpace hardware-in-the-loop system. In order to enable the code to run fast enough the
linearized mechanical model described in Sec. 3.4 was utilized. The table is now suc-
cessfully running with the controller to the defined 120 Hz closed-loop bandwidth spec-
ification. Note that the controller design is based on a linear model, requiring all the
mechanical parameters discussed in Sec. 3.4 to be known, along with the parameters of
linearized hydraulic Eqs. (A27) and (A28), which also require the valve response param-
eters of Eq. (17).
Another use of a detailed model of this type is in performance prediction of new sys-
tems to inform the design process and help in component selection. In addition, the
response prediction for an existing system with a particular specimen and test requirement
may highlight any performance issues before errors are made in testing a costly or unique
specimen. Ultimately, it may be possible to use a model to help pre-shape the command
signal to improve accuracy in the real test; however, this requires that the sensitivity of
system response to unknown specimen dynamics is small.

Acknowledgments
The author wishes to thank Instron Structural Testing Systems (IST Gmbh), Darms-
tadt, and IWHR, Beijing, for permission to publish this paper. In particular, the author
would like to thank Manfred Dickmann of IST for collecting the experimental data.
Most of the work described was undertaken when the author was employed by
Instron.
Detailed Dynamic Model of a Six-Axis Shaking Table 655

References
Chase, J. G, Hudson, N. H., Lin, J., Elliot, R., and Sim, A. [2005] “Non-linear shake table identifica-
tion and control for near-field earthquake testing,” Journal of Earthquake Engineering 9(4),
461–482.
Conte, J. P., and Trombetti, T. L. [2000] “Linear dynamic modelling of a uni-axial servo-
hydraulic shaking table system,” Earthquake Engineering and Structural Dynamics 29,
1375–1404.
Iwasaki, M., Ito, K., Kawafuku, M., Hirai, H., Dozono, Y., and Kuosaki, K. [2005] “Disturbance
observer based practical control of shaking tables with non-linear specimen,” 16th IFAC World
Congress, Prague.
Kakegawa, T., Suzuki, T., Sato, E., Kajiwara, K., and Tagawa, Y. [2003] “Linear model deriva-
tion and three-variable control (TVC) of the world's largest 3-D full-scale shaking table,”
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Transactions of the Japan Society of Mechanical Engineers, Part C, 69(2), 343–348 (in
Japanese).
Kuehn, J., Epp, D., and Patten, W. N. [1999] “High-fidelity control of a seismic shake table,” Earth-
quake Engineering and Structural Dynamics, 28, 1235–1254.
McCloy, D., and Martin, H. R. [1980] Control of Fluid Power: Analysis and Design. Ellis Horwood,
Chichester.
Merritt, H. E. [1967] Hydraulic Control Systems. John Wiley & Sons, New York.
Ogawa, N., Ohtani, K., Katayama, T., and Shibata H. [2001] “Construction of a three-dimensional,
large-scale shaking-table and development of core technology,” Philosophical Transactions
Rural of the Society of London A, 359, 1725–175.
Plummer, A. R., [2001] “Iterative velocity control for a high speed hydraulic actuator,” PTMC 2001,
Bath, September, 339–354.
Plummer, A. R. [2004] “Modal control for a class of multi-axis vibration table” Control 2004, Bath,
UK, September.
Plummer, A. R. [2007] “Control techniques for structural testing: a review” Proceedings of the
Instn. Mechanical Engineers, Part I, Journal of Systems and Control Engineering, 221 (2),
139–169.
Sato, E., Suzuki, T., Tagawa, Y., Kakegawa, T., Kajiwara, K., and Takai, S. [2002] “Alternative
control design approach to shaking table facilities for re-creating seismic motion,” Proceedings
of the ASME Pressure Vessels and Piping Conference, Vancouver, August, 67–73.
Shimizu, N., Shinohara, Y., and Sato, E. [2002] “New development of control method of shaking
table with bi-linear structures” Proceedings of the ASME Pressure Vessels and Piping Confer-
ence, Vancouver, August, 75–82.
Shimizu, N., Shinohara, Y., Yabuki, H., and Sato, E. [2004] “Control simulation of shaking table
with non-linear structure,” Proceedings of ACMD2004, 478–485.
Shortreed, J. S., Seible, F., and Benzoni, G., [2002] “Simulation issues with a real-time, full-scale
seismic testing system,” Journal of Earthquake Engineering 6(1), 185–201.
Trombetti, T. L. and Conte, J. P. [2002] “Shaking table dynamics: results from a test-analysis com-
parison study,” Journal of Earthquake Engineering 6(4), 513–551.

Appendix 1: Parameter Estimation

Structural Stiffness and Damping Parameters


Mounting and pushrod/joint stiffness and damping are found from the piston acceleration
over piston-to-cylinder frequency response. The lengths Ya, Yc, Ym, Yp, and Yt in Fig. 23
are for the steady-state zero force condition, and changes in these lengths are denoted ya,
yc, ym, yp, and yt, respectively. The mass of the cylinder body and piston rod are very small
compared to the table mass and they will be neglected. Let f be the compressive force
exerted by the cylinder, giving:
656 A. R. Plummer

Km Kp

Table
(mass Mt)
Cm Cp
Ym Yc Yp
Ya
Yt

FIGURE 23 Structural compliance modeling.


Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

yt = f
Mt && (A1)

(Cm s + K m ) ym = − f (A2)

(C p s + K p ) y p = − f . (A3)

The displacement of the table is the following sum of displacements:

yt = ya + y p (A4)

yt = ym + yc + y p . (A5)

Substituting (A1) and (A3) into (A4) gives:

⎛ Mt s2 ⎞
ya = ⎜ 1 + ⎟ yt . (A6)
⎝ Cps + K p ⎠

Substituting (A1) to (A3) into (A5) gives:

⎛ Mt s2 Mt s2 ⎞
yc = ⎜ 1 + + ⎟ yt . (A7)
⎝ Cm s + K m C p s + K p ⎠

A relationship between the absolute piston acceleration and the relative piston-to-cylinder
body displacement can be found from Eqs. (A6) and (A7):

⎛ Mt s2 ⎞ ⎛ Mt s2 Mt s2 ⎞
s2 ⎜ 1 + ⎟ yc = ⎜ 1 + + ⎟ &&
ya (A8)
⎝ Cps + K p ⎠ ⎝ Cm s + K m C p s + K p ⎠
Detailed Dynamic Model of a Six-Axis Shaking Table 657

⎛ Mt s2 (C p s + K p ) ⎞
s2 ( Mt s2 + C p s + K p ) yc = ⎜ + Mt s2 + C p s + K p ⎟ &&
ya (A9)
⎝ Cm s + K m ⎠

⎛ Cp K p s + 1⎞ K p 2 Cp Kp 2
s2 (s2 yc − &&
ya ) = ⎜ ⎟ ya −
s && ya ) −
s(s2 yc − && (s yc − &&
ya ). (A10)
⎝ C m K m s + 1⎠ K M Mt Mt

The requirement is to estimate the unknown stiffness and damping parameters from the
experimental frequency response of piston acceleration over relative actuator displace-
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

ment, i.e., from:

− w 2Ya ( j w )
= G( j w ) = gr (w ) + jgi (w ). (A11)
Yc ( j w )

A linear-in-the-parameters equation can be derived from (A10) by making an initial guess


that Cm has a value such that:

Cm K m = C p K p = t . (A12)

A better value for Cm could be found using an iterative procedure based on repeating the
estimation method described below, but in fact the dynamic behavior is not very sensitive
to Cm and so this has been found to be unnecessary.
Using Eq. (A12), (A10) becomes:

s2 (s2 yc − &&
ya ) = as2 &&
ya − bs(s2 yc − &&
ya ) − c(s2 yc − &&
ya ) (A13)

where a = Kp/Km, b = Cp/Mt, and c = Kp/Mt. The parameters to be estimated are a, b, and c,
from which all unknown stiffness and damping values can be found. In the frequency
domain, (A13) can be rewritten as:

w 2 (w 2 + G( j w )) = − aw 2 G( j w ) + jbw (w 2 + G( j w )) + c(w 2 + G( j w )) (A14)

or

⎡a ⎤
⎡ w 2 (w 2 + gr (w )) ⎤ ⎡ − w 2 gr (w ) − w gi (w ) w 2 + gr (w ) ⎤ ⎢ ⎥
⎢ ⎥=⎢ ⎥ b .
gi (w ) ⎥⎦ ⎢⎢ ⎥⎥
(A15)
⎢⎣ w 2 gi (w ) ⎥⎦ ⎢⎣ − w 2 gi (w ) w (w 2 + gr (w ))
⎣c ⎦

A least-squares fit gives the estimates for parameters a, b, and c.


658 A. R. Plummer

Hydraulic Parameter Values


The model equations presented in Sec. 3 can be linearized. The linear model will be a rea-
sonable approximation for an actuator operating near mid-stroke and with fairly low flows
and low differential pressure (p1 – p2). Given these limitations, this (or a model linearised
around any other operating point) should be used with care.
Using the following approximation, which is exact at mid-stroke,

v1 ≈ v2 ≈ V (A16)

each side of the system is the same (symmetrical) and it can be shown that
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

q1 = q2 = q (A17)

and

pa − pv1 = pv 2 − Pr for xv ≥ 0 (A18)

pa − pv 2 = pv1 − Pr for xv < 0. (A19)

Thus, adding Eqs. (1) and (2) and dividing by 2 gives:

V
q = Ay&c + (p&1 − p& 2 )+ Cl ( p1 − p2 ) + Cb f ( p1 − p2 ). (A20)
2B

Linearizing the cross-port bleed term about a nominal pressure differential ΔP:

Cb
Cb f( p1 − p2 ) ≈ ( p1 − p2 ), (A21)
2 ΔP

Eq. (A20) can be written as

q ⎛ 1 1⎞
= syc + ⎜ s + ⎟ fh (A22)
A ⎝ Kh Ch ⎠

where

2 BA2
Kh = (A23)
V

and

A2
Ch = . (A24)
Cl + Cb /(2 ΔP )
Detailed Dynamic Model of a Six-Axis Shaking Table 659

Linearizing the valve flow Eqs. (10–13) assuming the maximum available pressure is
dropped across the valve, and neglecting valve overlap:

⎛ P − Pr ⎞
q = ⎜ Kv s ⎟x (A25)
⎝ 2 ⎠

or

⎛ P − Pr ⎞
q = ⎜ Kv s V (s)u. (A26)
⎝ 2 ⎟⎠
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Finally, Eq. (A22) can be rewritten as

⎛ 1 1 ⎞
yc = yq − ⎜ + f (A27)
⎝ K h Ch s ⎟⎠

where yq is defined by

⎛K Ps − Pr ⎞ V (s)
yq = ⎜ v u. (A28)
⎝ A 2 ⎟⎠ s

Based on Eq. (A27), the linearized actuator model can be conceptualised as the stiffness of
the oil trapped in the cylinder, Kh, a series damper due to the leakage and cross-port bleed,
Ch, and the ‘ideal’ displacement yq, given purely by the flow of oil from the valve, i.e., this
would be the actuator displacement if there were no leakage and the oil were incompress-
ible. This concept is included in Fig. 24.
Substituting (A1–A3) into (A5) gives:

⎛ 1 1 1 ⎞
yc = ⎜ + + ⎟ f. (A29)
⎝ Mt s
2 Cm s + K m C p s + K p ⎠

Km Ch Kh Kp

yq Table
(mass Mt)
Cm Cp
Ym Yc Yp
Ya
Yt

FIGURE 24 Representation of linearized actuator model.


660 A. R. Plummer

Thus, making the same assumption about the damping values as in Eq. (A12) gives

f = F (s) yc (A30)

where

Mt s2 (t s + 1)
F ( s) = .
⎛ 1 1 ⎞ (A31)
⎜ K + K ⎟ Mt s + t s + 1
2
⎝ m p⎠
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

Thus from (A27):

V ( s) ⎛ e⎞
yc = g u + ⎜ d + ⎟ F (s) yc (A32)
s ⎝ s⎠

where

Kv Ps − Pr 1 1
g= ,d= and e = . (A33)
A 2 Kh Ch

The transfer functions V(s) and F(s) are known: V(s) is the valve spool dynamics
which, when the spool position is measured, can be estimated a priori; F(s) depends
on the mass and mechanical stiffnesses parameters estimated previously. Thus g, d,
and e are the parameters to be estimated. The following frequency response is
measured:

Yc ( jω )
= H ( jω ) = hr (ω ) + jhi (ω ). (A34)
U ( jω )

Thus, Eq. (A32) can be transformed to the frequency domain to give:

V ( jw ) ⎛ e⎞
H ( jw ) = g + ⎜ d + ⎟ F ( j w ) H ( j w ). (A35)
jw ⎝ jw ⎠

Let

V ( jw )
= vr (w ) + jvi (w ) (A36)
jw

F ( j w ) H ( j w ) = nr (w ) + jni (w ). (A37)
Detailed Dynamic Model of a Six-Axis Shaking Table 661

Thus, (A35) can be written as:

⎡ ni (w ) ⎤ g
⎡ ⎤
⎢ vr (w ) nr (w )
⎡ r ⎤
h ( w ) w ⎥⎢ ⎥
⎢ h (w ) ⎥ = ⎢ ⎥ ⎢d ⎥ .
(A38)
⎣ i ⎦ ⎢ v (w ) n (w ) − nr (w ) ⎥ ⎢ e ⎥
⎢⎣ i w ⎥⎦ ⎣ ⎦
i

A least-squares fit gives the estimates for parameters g, d, and e.

Appendix 2 – Notation
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

A piston face area


B bulk modulus
Cb cross-port bleed
Ch internal actuator damping
Cl internal cylinder leakage
Cm cylinder mounting damping
Cp pushrod/joint damping
D valve response: dead time
F(s) force over piston-to-cylinder displacement transfer function
f force on table or hydraulic force (neglecting piston mass)
fh hydraulic force (product of piston area and differential pressure)
ft or fti force on table from individual actuator (actuator i)
fx, fy, fz forces on table in orthogonal Cartesian co-ordinates
ft vector of N actuator forces on table
ft vector of six resultant forces and moments in Cartesian co-ords
H(s) open-loop transfer function
Ixx, Iyy, Izz table moments of inertia (including rigid specimen where present)
Ixy, Ixz, Iyz table products of inertia (including rigid specimen where present)
J Jacobean matrix
Kb valve body flow coefficient (body saturation)
Kf manifold flow coeffcient
Kh internal actuator oil-column stiffness
Km cylinder mounting stiffness
Kp pushrod/joint stiffness
Kv valve flow coefficient
L valve overlap
Lh offset of x-direction actuators from center line
Lv offset of z-direction actuators from center line
M table mass matrix
Mc cylinder body mass
Mp piston mass
Mt table mass including specimen (where present)
N number of actuators
n polytropic index
P0 accumulator precharge pressure
Pr return pressure
Ps supply pressure
662 A. R. Plummer

ΔP pressure difference
p1 pressure on side 1 of piston
p2 pressure on side 2 of piston
pv1 pressure at cylinder side of valve orifice (side 1)
pv2 pressure at cylinder side of valve orifice (side 2)
pa accumulator pressure
Qp available pump flow rate
q volume flow rate
q1 volume flow rate into side 1 of cylinder
q2 volume flow rate out of side 2 of cylinder
Qp power pack flow (after loss through relief valve)
qsi supply flow rate required for actuator i
Downloaded by [Eindhoven Technical University] at 15:26 06 January 2015

s differential operator
Sr valve maximum spool velocity (slew rate limit)
u control signal
V oil volume one side of piston when at mid-stroke
V(s) valve dynamics model
V0 accumulator volume
Vs accumulator gas volume at pressure Ps
v1 oil volume on side 1 of piston
v2 oil volume on side 2 of piston
x normalized valve spool position
xv valve orifice opening
Ya length: cylinder mount to piston rod joint (mid-stroke, steady state)
Yc length: piston rod beyond cylinder body (mid-stroke, steady state)
Ym length: cylinder body (steady state)
Yp length: pushrod (steady state)
Yt length: cylinder mount to table (mid-stroke, steady state)
ya deviation from Ya during dynamic operation
yc deviation from Yc during dynamic operation
ym deviation from Ym during dynamic operation
yp deviation from Yp during dynamic operation
yt deviation from Yt during dynamic operation
yx , yy , yz table linear displacements in Cartesian co-ordinates
yt vector of N table displacements at actuator attachment points
yt vector of six linear and angular displacements in Cartesian co-ords
f(•) function of pressure used for orifice equation
jx, jy, jz table angular displacements in Cartesian co-ordinates
t damping over stiffness ratio
w frequency
wnv valve natural frequency
zv valve damping ratio

Вам также может понравиться